Next Article in Journal
Raman Identification of Polymorphs in Pentacene Films
Previous Article in Journal
High-Spin Chains and Crowns from Double-Exchange Mechanism
Previous Article in Special Issue
Preparation, Crystal and Properties of Nitrogen-Rich Energetic Salt of Bis(semicarbazide) 5,5′-Bitetrazole-1,1′-diolate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

2-Azidoimidazolium Ions Captured by N-Heterocyclic Carbenes: Azole-Substituted Triazatrimethine Cyanines

1
Faculty of Chemistry and Pharmacy, University of Innsbruck, Innsbruck 6020, Austria
2
Institute of Mineralogy and Petrography, University of Innsbruck, Innsbruck 6020, Austria
3
Research Institute for Textile Chemistry and Textile Physics, University of Innsbruck, Dornbirn 6850, Austria
*
Author to whom correspondence should be addressed.
Crystals 2016, 6(4), 40; https://doi.org/10.3390/cryst6040040
Submission received: 29 January 2016 / Revised: 15 February 2016 / Accepted: 17 February 2016 / Published: 8 April 2016
(This article belongs to the Special Issue Nitrogen-Rich Salts)

Abstract

:
1,3-Disubstituted 2-azidoimidazolium salts (substituents = methyl, methoxy; anion = PF6) reacted with N-heterocyclic carbenes to yield yellow 2-(1-(azolinylidene)triazen-3-yl)-1,3-R2-imidazolium salts (azole = 1,3-dimethylimidazole, 1,3-dimethoxyimidazole, 4-dimethylamino-1-methyl-1,2,4-triazole; R = methyl, methoxy; anion = PF6). Crystal structures of three cationic triazenes were determined. Numerous interionic C–H···F contacts were observed. Solvatochromism of the triazenes in polar solvents was investigated by UV-Vis spectroscopy, involving the dipolarity π* and hydrogen-bond donor acidity α of the solvent. Cyclovoltammetry showed irreversible reduction of the cations to uncharged radicals. Thermoanalysis showed exothermal decomposition.

Graphical Abstract

1. Introduction

Previously, the 2-azido-1,3-dimethoxyimidazolium [1] and 2-azido-1,3-dimethylimidazolinium [2] ions were synthesized from the respective 2-bromo or 2-chloro compounds and sodium azide. This pathway is only viable when the halogen derivatives are readily available. A different approach to azidoazoles involves fragmentation of sulfonyltriazenes [3], which can be accessed by azidation of azole anions. In addition, triazene formation is accomplished by reaction of imidazolin-2-ylidenes with alkyl or aryl azides [4,5], and a series of triazene crystal structures of this type are known [4,5,6,7,8,9,10,11]. One crystal structure of a cationic triazatrimethine cyanine chromophore derived from benzimidazole has been reported [12]. Other related structures include metal complexes of neutral [13,14,15] and anionic 1,3-bis(tetrazol-5-yl)triazene ligands [16,17]. In this work, we intended to combine the azidation of a carbene with the fragmentation of a sulfonyltriazene.

2. Results and Discussion

Since the 2-azido-1,3-dimethylimidazolium ion is not yet described, we attempted to prepare it by azidation of the respective carbene using tosyl azide (Figure 1).
Surprisingly, it was found that the resulting tosyltriazene reacted with the intermediate carbene, yielding the disubstituted triazene 1. The primary product mixture consisted of lithium toluenesulfinate and triazene and was easily separated. The desired 2-azidoimidazolium salt could not be isolated but will be attempted in further work. However, this result inspired us to prepare other triazenes, or triazatrimethine cyanines, from stable azidoazolium salts and N-heterocyclic carbenes. Thus, several methods to generate the carbenes were tested, and sodium hydride in acetonitrile was discovered to give agreeable results and allowed the synthesis of another symmetrically substituted triazene 2 and the unsymmetrically substituted triazene 3.
The hexafluoridophosphate anions were chosen because they readily yield non-hygroscopic, crystalline solids. The role of fluorine, which exhibits the lowest polarizability and highest electronegativity of the halogens in crystal structures of organic compounds has been reviewed [18]. After some controversy, it has been concluded that short C–H···F contacts between oppositely charged molecules are genuine interionic hydrogen bonds [19].
The crystal structures of the new cationic triazatrimethine cyanines 13 were determined and are discussed below.

2.1. Crystal Structures

The C10H16N7 cation in triazene 1 contains 41.9% N and may be termed “nitrogen-rich”. A crystal of 1 was refined as a two-component twin. Positional disorder of the fluorine atoms of PF6 was observed, and a ratio of 3:2 was obtained by refinement as a free variable. In addition, the small crystal was non-merohedrally twinned by 180 degrees about the reciprocal axis 0 0 1.
The imidazole-substituted triazene system in 1 is heavily distorted. The angle between the planes of the two imidazole rings is 59.3°, the N5–N6 bond is rotated out of the plane of the first imidazole ring by 25.9°, and the N7–N6 bond is rotated out of the plane of the second imidazole ring by 32.4° (Figure 2a). Although the mesomeric cyanine molecule is symmetrically disubstituted, the bond lengths in the triazene bridge do not reflect this symmetry, presumably due to packing effects.
The heteroaromatic C2–H, C7–H and C8–H donate hydrogen bonds to F4, F3 and F1, whereas H atoms of one methyl group form contacts to F1 and F6 (Figure 2b).
Crystal data and structure refinement details are summarized in Table 1. Short contacts and interactions are listed in Table 2. Selected bond lengths, angles and torsions of the triazene bridge are compiled in Table 3.
The imidazole-substituted triazene system in 2 is far less distorted than in 1. Thus, the angle between the planes of the two imidazole rings is 14.8°, the N5–N6 bond is twisted out of the plane of the adjacent imidazole ring by 7.6°, and the N7–N6 bond by 16.3°. The methoxy groups of both imidazole rings adopt syn conformations, although in opposite directions (Figure 3a).
Several C–H···F interactions involve the heteroaromatic C3–H, C7–H and C8–H, as well as the C9 methyl group. Additionally, C2–H···O3 and C4–H···O4 interactions are observed (Figure 3b).
The PF6-groups in 2 and 3 were orientationally disordered among two positions. SADI commands were employed to restrain the corresponding P–F and F–F distances. For the refinement of the anisotropic displacement parameters, SIMU commands were used for atoms closer than 1.0 Å.
The unsymmetrically disubstituted triazene system in 3 is the least distorted of the three compounds. The angle between the planes of the two heterocyclic rings is only 4.8°, the N5–N6 bond is rotated out of the imidazole plane by 10.5°, and the N7–N6 bond out of the triazole plane by 13.5° (Figure 4a). Again, the methoxy groups adopt syn conformation. In contrast to 1, although the cyanine molecule is unsymmetrically disubstituted, the corresponding bond lengths in the triazene bridge are almost equal. A comparison of the bond lengths in the triazenes with accepted values [20] reveals the delocalized nature of the bonds in these systems. The N···N distances here are longer than trans N=N bonds (1.22 A), but shorter than typical planar N–N bonds (1.40 A). In addition, the C···N distances are longer than C=N bonds (1.28 A) but shorter than isolated C–N= bonds (1.43 A). The bonds in the triazenes compare favorably to aromatic N N (1.30 A) and C N (1.34 A) bonds.
In the structure of 3, the heteroaromatic C7–H of the triazole donates a hydrogen bond to F5, whereas the C3–H of the imidazole forms bifurcated hydrogen bonds to F3 and F6. Aliphatic hydrogen atoms form contacts to F1, F3 and F5, as well as to O1 and N7 (Figure 4b).

2.2. UV-Vis Spectroscopy

Typically, donor–acceptor-substituted conjugated systems exhibit solvatochromism. The solvatochromism of such systems arises from the different stabilization of their electronic ground and excited state by differential solvation of these two states, according to their different molecular and electronic structure. Transition energies ET were calculated from the wavelengths of the absorption maxima (Table 4) according to the equation ET/kJ·mol−1 = hcN/λ = 119625/(λ/nm). A linear solvation energy relationship was established by least-squares fitting the data to the solvatochromic equation involving the parameters π* [21] and α [22] which represent the dipolarity/polarizability and hydrogen-bond donor capabilities of the solvent.The hydrogen-bond acceptor basicity β [23] did not significantly contribute to the relationship. The following equations were derived by multiple linear regression (Figure 5), which adequately describe the solvatochromic behavior of the triazenes.
For 1: ET/kJ·mol−1 = 291.7 + 3.98 π* + 8.61 α (n = 7, R2 = 0.992)
For 2: ET/kJ·mol−1 = 313.2 − 9.21 π* − 5.08 α (n = 6, R2 = 0.960)
For 3: ET/kJ·mol−1 = 317.96 + 3.32 π* + 8.15 α (n = 6, R2 = 0.960)
The triazenes 1 and 3 exhibited negative solvatochromism (shift of the absorption maximum to shorter wavelengths on changing the solvent from dichloromethane to water) whereas, amazingly, triazene 2 displayed positive solvatochromism. No correlation was found with less polar solvents.
The values in DMSO for 2 and 3 were not included in the calculation because a change of band shape rendered them not comparable. The π* and α values were taken from the compilation of Marcus [24].

2.3. Differential Scanning Calorimetry (DSC) and Thermogravimetric Analysis (TGA)

Thermoanalysis showed exothermal decomposition (maxima at 282 for 1, 180 for 2 and 186 °C for 3) with considerable mass loss (Figure 6).

2.4. Cyclic Voltammetry (CV)

The three electrophoric systems have in common that no reverse peaks were observed. Presumably, the first electrochemical process involved an irreversible reduction to an uncharged radical species. The resulting radical evidently was unstable on the electrochemical time scale and underwent subsequent decomposition. The cathodic peak current (Ip)c was approximately proportional to the square root of the scan rate, and the cathodic peak potential (Ep)c shifted to negative values with an increasing rate (Figure 7). The peak potential and peak current values found in the voltammetric measurements are summarized in Table 5.

3. Experimental Section

1,3-Dimethylimidazolium hexafluoridophosphate, 2-azido-1,3-dimethoxyimidazolium hexafluoridophosphate, 1,3-dimethoxyimidazolium hexafluoridophosphate, and 4-(dimethylamino)-1-methyl-1,2,4-triazolium hexafluoridophosphate were prepared as described previously [1,25,26]. All other chemicals were purchased from Sigma-Aldrich, St. Louis, MO, USA (European affiliate, Steinheim, Germany). NMR spectra were recorded with a Bruker Avance DPX 300 spectrometer (Billerica, MA, USA). IR spectra were obtained with a Nicolet 5700 FT spectrometer (Thermo Fisher Scientific Inc., Waltham, MA, USA) in ATR mode. UV-Vis spectra were recorded with a Perkin-Elmer Lambda XLS+ spectrometer (Waltham, MA, USA); mean values of at least three replicates were taken. High resolution mass spectra were measured with a Finnigan MAT 95 mass spectrometer (Waltham, MA, USA). DSC and TGA were recorded with Perkin-Elmer DSC 7 and TGA 7 instruments (Waltham, MA, USA) at a heating rate of 10 °C·min−1 under nitrogen.
Cyclic voltammetric (CV) experiments were performed in CH3CN at room temperature under Ar atmosphere using an EG&G 264 A polarographic analyser/stripping voltammeter (Princeton Applied Research, Oak Ridge, TN, USA). All measurements were carried out employing a three-electrode configuration consisting of a Pt electrode (disk, 2 mm diameter) as a working electrode and an (Ag/0.01 M AgNO3 in CH3CN) electrode as the reference (EG&G Micro-Cell). A Pt wire in an electrode bridge tube filled with supporting electrolyte and separated from the sample by a porous frit served as the counter electrode. A solution of 0.1 M Bu4PF6 in CH3CN served as the supporting electrolyte. Sample solutions were prepared by dissolving triazene (2.5 mg) in electrolyte (25 mL). A solution of ferrocene in electrolyte was used for calibration. All formal potentials are referenced to the ferrocene/ferrocenium redox couple.
Single crystal diffraction intensity data were recorded by φ and ω scans with a Bruker D8 Quest Photon 100 (Billerica, MA, USA) (compound 1), or by ω scans with an Oxford Diffraction Gemini-R Ultra (Oxford Diffraction Ltd., Abingdon, Oxfordshire, UK) (compounds 2 and 3) diffractometer using MoKα radiation. The structures were solved by direct methods and refined by full-matrix least-squares techniques on F2. Programs CELL_Now and TWINABS (Bruker) were used for cell search of twin components and absorption correction. CCDC 1428586–1428588 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html.

3.1. 1,3-Dimethyl-2-(1-(1,3-dimethylimidazolin-2-ylidene)triazen-3-yl)imidazolium Hexafluoridophosphate (1)

A solution of lithium bis(trimethylsilyl)amide in THF (5.4 mL 1 M, 1.3 equiv) was added to a suspension of 1,3-dimethylimidazolium hexafluoridophosphate (1.0 g, 4.1 mmol) in THF (10 mL) at 0 °C. The resulting solution was stirred for 30 min at 0 °C, then tosyl azide (0.41 g, 0.5 equiv) in THF (1 mL) was added. The mixture was stirred for 4 h at room temperature, then cooled at 0 °C again. The orange precipitate was collected by filtration and partitioned between CH2Cl2 (15 mL) and H2O (15 mL). The aqueous phase was extracted once more with CH2Cl2 (10 mL), and the combined organic phases were dried over MgSO4 and taken to dryness under reduced pressure to yield the triazene 1 as an orange powder (0.25 g, 35%). Single crystals were obtained from H2O/acetone. M.p. 201 °C (decomposition). 1H NMR (DMSO-d6, 300 MHz, δ): 3.68 (s, 12H), 7.39 (s, 4H) ppm. 13C NMR (DMSO-d6, 75 MHz, δ): 35.1 (4C), 119.8 (4C), 147.9 (2C) ppm. IR (neat): ν ˜ 3174 (w), 3143 (w), 1573 (w), 1520 (m), 1503 (m), 1343 (w), 1272 (m), 1222 (m), 1099 (m), 827 (s), 755 (m), 711 (m), 700 (m), 555 (s) cm−1. HRMS (ES): m/z = 234.143 (calcd. 234.146 for C10H16N7, [M]+).

3.2. 1,3-Dimethoxy-2-(1-(1,3-dimethoxyimidazolin-2-ylidene)triazen-3-yl)imidazolium Hexafluoridophosphate (2)

1,3-Dimethoxyimidazolium hexafluoridophosphate (0.10 g, 0.0004 mol) and 2-azido-1,3-dimethoxyimidazolium hexafluoridophosphate (0.12 g, 0.0004 mol) were dissolved in CH3CN (5 mL). The resulting solution was cooled to 0 °C, and NaH (0.015 g, 0.0004 mol) was slowly added. A yellow precipitate was formed instantaneously, and the mixture was stirred for another 3 h. Subsequently, the solvent was evaporated, and the yellow residue was recrystallized from H2O/acetone leading to yellow plates. Yield: 0.04 g (25%). M.p. 165 °C (decomposition). 1H NMR (DMSO-d6, 300 MHz, δ): 4.08 (s, 12H), 7.98 (s, 4H) ppm. 13C NMR (DMSO-d6, 75 MHz, δ): 67.9 (4C), 113.5 (4C), 141.5 (2C) ppm. IR (neat): ν ˜ 3167 (w), 3147 (w), 1556 (m), 1500 (m), 1247 (m), 1159 (m), 1053 (m), 1044 (m), 946 (m), 824 (s), 677 (m), 554 (s) cm−1. HRMS (ES): m/z = 298.124 (calcd. 298.126 for C10H16N7O4, [M]+).

3.3. 1,3-Dimethoxy-2-(1-(1-methyl-4-dimethylamino-1,2,4-triazolin-5-ylidene)triazen-3-yl)imidazolium Hexafluoridophosphate (3)

4-(Dimethylamino)-1-methyltriazolium hexafluoridophosphate (0.31 g, 0.0011 mol) and 2-azido-1,3-dimethoxyimidazolium hexafluoridophosphate (0.36 g, 0.0011 mol) were dissolved in CH3CN (5 mL). The resulting solution was cooled to 0 °C, and NaH (0.05 g, 0.0011 mol) was slowly added. A yellow precipitate was formed instantaneously, and the mixture was stirred for another hour. Subsequently, the solvent was evaporated, and the residue was washed with H2O (3 × 2 mL) and acetone (2 × 2 mL) leading to a yellow powder. The crude product was recrystallized from H2O/acetone yielding bright yellow needles. Yield: 0.22 g (43%). M.p. 172 °C (decomposition). 1H NMR (DMSO-d6, 300 MHz, δ): 2.95 (s, 6H), 3.91 (s, 3H), 4.10 (s, 6H), 8.03 (s, 2H), 9.25 (s, 1H) ppm. 13C NMR (DMSO-d6, 75 MHz, δ): 46.3 (2C), 68.0 (2C), 113.7 (2C), 139.7, 141.1, 151.8 ppm. 1H NMR (acetone-d6, 300 MHz, δ): 3.10 (s, 6H), 4.04 (s, 3H), 4.23 (s, 6H), 7.81 (s, 2H), 8.84 (s, 1H) ppm. 13C NMR (acetone-d6, 75 MHz, δ): 40.3, 46.7 (2C), 68.8 (2C), 114.6 (2C), 140.7 (2C), 154.0 ppm. IR (neat): ν ˜ 3167 (w), 3152 (w), 2923 (w), 2851 (w), 1546 (m), 1506 (m), 1453 (m), 1272 (m), 1242 (m), 1170 (m), 1160 (m), 1047 (m), 817 (s), 682 (m), 555 (s) cm−1. HRMS (ES): m/z = 298.155 (calcd. 296.158 for C10H18N9O2, [M]+).

4. Conclusions

In the growing field of N-heterocyclic carbene (NHC) catalysis, one of the most recent and exciting concepts is the thermally triggered release of the catalytically active species [27]. The new azole-substituted triazenes could become of interest as metal-free, labile NHC progenitors capable of switching on latent reactions. More research is needed with regard to solvent and temperature applied in order to liberate an active NHC.

Acknowledgments

The authors are grateful to Ulrich J. Griesser and Elisabeth Gstrein for the generous measurements of DSC and TGA.

Author Contributions

Simone Haslinger carried out experimental work (synthesis, crystallization and characterization), and Gerhard Laus and Herwig Schottenberger conceived and designed this study. Gerhard Laus wrote the manuscript. Volker Kahlenberg and Klaus Wurst determined the crystal structures. Thomas Bechtold contributed the cyclic voltammetry measurements. Stefan Vergeiner recorded the mass spectra.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Laus, G.; Schwärzler, A.; Schuster, P.; Bentivoglio, G.; Hummel, M.; Wurst, K.; Kahlenberg, V.; Lörting, T.; Schütz, J.; Peringer, P.; et al. N,N′-Di(alkyloxy)imidazolium Salts: New Patent-free Ionic Liquids and NHC Precatalysts. Z. Naturforsch. 2007, 62b, 295–308. [Google Scholar] [CrossRef]
  2. Kitamura, M.; Kato, S.; Yano, M.; Tashiro, N.; Shiratake, Y.; Sando, M.; Okauchi, T. A reagent for safe and efficient diazo–transfer to primary amines: 2–azido–1,3–dimethylimidazolinium hexafluorophosphate. Org. Biomol. Chem. 2014, 12, 4397–4406. [Google Scholar] [CrossRef] [PubMed]
  3. Zanirato, P.; Cerini, S. On the utility of the azido transfer protocol: Synthesis of 2- and 5-azido N-methylimidazoles, 1,3-thiazoles and N-methylpyrazole and their conversion to triazole–azole bisheteroaryls. Org. Biomol. Chem. 2005, 3, 1508–1513. [Google Scholar] [CrossRef] [PubMed]
  4. Khramov, D.M.; Bielawski, C.W. Triazene formation via reaction of imidazol-2-ylidenes with azides. Chem. Commun. 2005, 39, 4958–4960. [Google Scholar] [CrossRef] [PubMed]
  5. Khramov, D.M.; Bielawski, C.W. Donor–Acceptor Triazenes: Synthesis, Characterization, and Study of Their Electronic and Thermal Properties. J. Org. Chem. 2007, 72, 9407–9417. [Google Scholar] [CrossRef] [PubMed]
  6. Kunetskiy, R.A.; Cisarova, I.; Saman, D.; Lyapkalo, I.M. New Lipophilic 2-Amino-N,N′-dialkyl-4,5-dimethylimidazolium Cations: Synthesis, Structure, Properties, and Outstanding Thermal Stability in Alkaline Media. Chem. Eur. J. 2009, 15, 9477–9485. [Google Scholar] [CrossRef] [PubMed]
  7. Tennyson, A.G.; Ono, R.J.; Hudnall, T.W.; Khramov, D.M.; Er, J.A.V.; Kamplain, J.W.; Lynch, V.M.; Sessler, J.L.; Bielawski, C.W. Quinobis(imidazolylidene): Synthesis and Study of an Electron-Configurable Bis(N-Heterocyclic Carbene) and Its Bimetallic Complexes. Chem. Eur. J. 2010, 16, 304–315. [Google Scholar] [CrossRef] [PubMed]
  8. Ghadwal, R.S.; Roesky, H.W.; Granitzka, M.; Stalke, D. A Facile Route to Functionalized N–Heterocyclic Carbenes (NHCs) with NHC Base-Stabilized Dichlorosilylene. J. Am. Chem. Soc. 2010, 132, 10018–10020. [Google Scholar] [CrossRef] [PubMed]
  9. Tennyson, A.G.; Moorhead, E.J.; Madison, B.L.; Er, J.A.V.; Lynch, V.M.; Bielawski, C.W. Methylation of Ylidene-Triazenes: Insight and Guidance for 1,3-Dipolar Cycloaddition Reactions. Eur. J. Org. Chem. 2010, 2010, 6277–6282. [Google Scholar] [CrossRef]
  10. Patil, S.; Bugarin, A. Crystal structure of (E)-1,3-dimethyl-2-[3-(3-nitrophenyl)triaz-2-en-1-ylidene]-2,3-dihydro-1H-imidazole. Acta Crystallogr. 2014, E70, 224–227. [Google Scholar] [CrossRef] [PubMed]
  11. Lysenko, S.; Daniliuc, C.G.; Jones, P.G.; Tamm, M. Tungsten alkylidyne complexes with ancillary imidazolin-2-iminato and imidazolidin-2-iminato ligands and their use in catalytic alkyne metathesis. J. Organomet. Chem. 2013, 744, 7–14. [Google Scholar] [CrossRef]
  12. Naef, R.; Balli, H. Synthesis, Structure and Photochemical Properties of 4,4′,7,7′-Tetra-substituted1,1′,3,3′-Tetraethylbenzimidazolotriazatrimethine Cyanines. Helv. Chim. Acta 1978, 61, 2958–2973. [Google Scholar] [CrossRef]
  13. Hanot, V.P.; Robert, T.D.; Kolnaar, J.J.A.; Haasnoot, J.G.; Kooijman, H.; Spek, A.L. Crystal structure and magnetic properties of a µ-halogeno-bridged copper(II) chain with neutral planar [Cu(batt)Cl] units (Hbatt = 1,3-bis[3-(5-amino-1,2,4-triazolyl)]triazene). Inorg. Chim. Acta 1997, 256, 327–329. [Google Scholar] [CrossRef]
  14. Hanot, V.P.; Robert, T.D.; Haasnoot, J.G.; Kooijman, H.; Spek, A.L. Crystal structure and resonance Raman spectroscopic study of {1,3-bis[3-(5-amino-1,2,4-triazolyl)]triazenido-N′4,N2,N″4}chloro palladium(II)-methanol (1/1). J. Chem. Cryst. 1998, 28, 343–351. [Google Scholar] [CrossRef]
  15. Hanot, V.P.; Robert, T.D.; Haasnoot, J.G.; Kooijman, H.; Spek, A.L. Crystal structure and spectroscopic study of bis{1,3-bis[3-(5-amino-1,2,4-triazolyl)]triazenido-N′4,N2,N″4}nickel(II) tetrahydrate. J. Chem. Cryst. 1999, 29, 299–308. [Google Scholar] [CrossRef]
  16. Serebryanskaya, T.V.; Ivashkevich, L.S.; Lyakhov, A.S.; Gaponik, P.N.; Ivashkevich, O.A. 1,3-Bis(2-alkyltetrazol-5-yl)triazenes and their Fe(II), Co(II) and Ni(II) complexes: Synthesis, spectroscopy, and thermal properties. Crystal structure of Fe(II) and Co(II) 1,3-bis(2-methyltetrazol-5-yl)triazenide complexes. Polyhedron 2010, 29, 2844–2850. [Google Scholar] [CrossRef]
  17. Eulgem, P.J.; Klein, A.; Maggiarosa, N.; Naumann, D.; Pohl, R.W.H. New Rare Earth Metal Complexes with Nitrogen-Rich Ligands: 5,5′-Bitetrazolate and 1,3-Bis(tetrazol-5-yl)triazenate–On the Borderline between Coordination and the Formation of Salt-Like Compounds. Chem. Eur. J. 2008, 14, 3727–3736. [Google Scholar] [CrossRef] [PubMed]
  18. Chopra, D.; Guru Row, T.N. Role of organic fluorine in crystal engineering. CrystEngComm 2011, 13, 2175–2186. [Google Scholar] [CrossRef]
  19. D’Oria, E.; Novoa, J.J. On the hydrogen bond nature of the C–H···F interactions in molecular crystals. An exhaustive investigation combining a crystallographic database search and ab initio theoretical calculations. CrystEngComm 2008, 10, 423–436. [Google Scholar] [CrossRef]
  20. Allen, F.H.; Kennard, O.; Watson, D.G.; Brammer, L.; Orpen, A.G.; Taylor, R. Tables of Bond Lengths determined by X-Ray and Neutron Diffraction. Part I. Bond Lengths in Organic Compounds. J. Chem. Soc. Perkin Trans. 1987, 2, S1–S19. [Google Scholar] [CrossRef]
  21. Kamlet, M.J.; Abboud, J.-L.M.; Taft, R.W. The Solvatochromic Comparison Method. The π* Scale of Solvent Polarities. J. Am. Chem. Soc. 1977, 99, 6027–6038. [Google Scholar] [CrossRef]
  22. Taft, R.W.; Kamlet, M.J. The Solvatochromic Comparison Method. The α Scale of Solvent Hydrogen-Bond Donor (HBD) Acidities. J. Am. Chem. Soc. 1976, 98, 2886–2894. [Google Scholar] [CrossRef]
  23. Kamlet, M.J.; Taft, R.W. The Solvatochromic Comparison Method. The β Scale of Solvent Hydrogen-Bond Acceptor (HBA) Basicities. J. Am. Chem. Soc. 1976, 98, 377–383. [Google Scholar] [CrossRef]
  24. Marcus, Y. The Properties of Organic Liquids that are Relevant to their Use as Solvating Solvents. Chem. Soc. Rev. 1993, 22, 409–416. [Google Scholar] [CrossRef]
  25. Holbrey, J.D.; Reichert, W.M.; Swatloski, R.P.; Broker, G.A.; Pitner, W.R.; Seddon, K.R.; Rogers, R.D. Efficient, halide free synthesis of new, low cost ionic liquids: 1,3-dialkylimidazolium salts containing methyl- and ethyl-sulfate anions. Green Chem. 2002, 4, 407–413. [Google Scholar] [CrossRef]
  26. Schwärzler, A.; Laus, G.; Kahlenberg, V.; Wurst, K.; Gelbrich, T.; Kreutz, C.; Kopacka, H.; Bonn, G.; Schottenberger, H. Quaternary 4-amino-1,2,4-triazolium Salts: Crystal structures of Ionic Liquids and N-heterocyclic carbene (NHC) complexes. Z. Naturforsch. 2009, 64b, 603–616. [Google Scholar] [CrossRef]
  27. Naumann, S.; Buchmeiser, M.R. Liberation of N-heterocyclic carbenes (NHCs) from thermally labile progenitors: Protected NHCs as versatile tools in organo- and polymerization catalysis. Catal. Sci. Technol. 2014, 4, 2466–2479. [Google Scholar] [CrossRef]
Figure 1. Synthesis of 13. Reagents and conditions: a: Li(Me3Si)2N, THF; b: TsN3; c: NaH, CH3CN.
Figure 1. Synthesis of 13. Reagents and conditions: a: Li(Me3Si)2N, THF; b: TsN3; c: NaH, CH3CN.
Crystals 06 00040 g001
Figure 2. (a) Molecular structure of the cation; (b) interactions in the crystal structure of 1.
Figure 2. (a) Molecular structure of the cation; (b) interactions in the crystal structure of 1.
Crystals 06 00040 g002
Figure 3. (a) Molecular structure of the cation; (b) interactions in the crystal structure of 2.
Figure 3. (a) Molecular structure of the cation; (b) interactions in the crystal structure of 2.
Crystals 06 00040 g003
Figure 4. (a) Molecular structure of the cation; (b) interactions in the crystal structure of 3.
Figure 4. (a) Molecular structure of the cation; (b) interactions in the crystal structure of 3.
Crystals 06 00040 g004
Figure 5. (a) Normalized UV-Vis spectra of triazenes 13 in H2O. (bd) correlation of observed vs. calculated transition energies of triazenes 13 in polar solvents.
Figure 5. (a) Normalized UV-Vis spectra of triazenes 13 in H2O. (bd) correlation of observed vs. calculated transition energies of triazenes 13 in polar solvents.
Crystals 06 00040 g005
Figure 6. (ac) DSC and TGA of the triazenes 13.
Figure 6. (ac) DSC and TGA of the triazenes 13.
Crystals 06 00040 g006
Figure 7. (ac) Cyclic voltammograms of the triazenes 13.
Figure 7. (ac) Cyclic voltammograms of the triazenes 13.
Crystals 06 00040 g007aCrystals 06 00040 g007b
Table 1. Crystal data and structure refinement for 13.
Table 1. Crystal data and structure refinement for 13.
Compound123
CCDC No.142858614285871428588
Empirical formulaC10H16N7·F6PC10H16N7O4·F6PC10H18N9O2·F6P
Formula weight379.27443.27441.3
Crystal systemTriclinicMonoclinicMonoclinic
Space groupP 1 ¯ P21/cP21
a7.608 (1)10.354 (1)10.567 (1)
b8.386 (1)13.153 (1)6.573 (1)
c13.806 (1)13.863 (1)14.267 (2)
α/°80.893 (2)
β/°77.396 (2)109.841 (8)111.390 (11)
γ/°70.006 (2)
Volume/Å3804.5 (1)1775.8 (2)922.6 (2)
Z242
Dx/g·cm−31.571.661.59
μ/mm−10.240.250.24
F (000)388904452
T/K170 (2)173 (2)173 (2)
Crystal size/mm30.18 × 0.15 × 0.030.40 × 0.28 × 0.080.38 × 0.32 × 0.28
θmax25.025.425.3
Index ranges–8 ≤ h ≤ 9,
–9 ≤ k ≤ 9,
0 ≤ l ≤ 16
–12 ≤ h ≤ 12,
–15 ≤ k ≤ 15,
–12 ≤ l ≤ 16
–12 ≤ h ≤ 10,
–7 ≤ k ≤ 7,
–15 ≤ l ≤ 17
Reflections collected3311112355556
Independent reflections (Rint)33113251 (0.026)3303 (0.023)
Observed reflections (I > 2σ (I))265127013215
Absorption correctionmulti-scanmulti-scanmulti-scan
Restraints/parameters0/278162/312157/313
Goodness-of-fit on F21.041.041.10
R1/wR2 (I > 2σ (I))0.057/0.1240.037/0.0880.031/0.078
R1/wR2 (all data)0.080/0.1340.048/0.0940.032/0.079
Δρmax/Δρmin/e·Å−30.277/–0.4260.44/–0.260.24/–0.19
Table 2. Interactions for 13 (Å, °).
Table 2. Interactions for 13 (Å, °).
CompoundInteractionH···AC···AC–H···ASymmetry Operation A
1C8–H···F12.4173.28 (1)151.0x, y, 1 + z
C9–HC···F12.4173.18 (1)134.51 − x, 1 − y, 1 − z
C7–H···F32.4223.26 (2)146.4x, −1 + y, 1 + z
C9–HA···F62.4562.99 (1)114.1x, 1 − y, 1 − z
C10–HA···F52.4633.19 (1)130.71 − x, 2 − y, 1 − z
C2–H···F42.5703.17 (1)131.21 + x, y, z
2C2–H···O32.2953.128 (3)146.01 − x, 1/2 + y, 3/2 − z
C7–H···F6A2.2813.21 (1)167.32 − x, −1/2 + y, 3/2 − z
C8–H···F3A2.3843.24 (1)149.82 − x, −y,1 − z
C9–HB···F1A2.4143.02 (1)119.5x, 1/2 − y, 1/2 + z
C3–H···F5A2.4603.30 (1)147.61 − x, 1 − y, 1 − z
C4–HC···O42.5753.355 (3)136.61 − x, −y, 1 − z
3C3–H···F6A2.5123.34 (1)145.3x, 1 + y, z
C3–H···F3A2.5423.27 (1)133.1x, 1 + y, z
C7–H···F5A2.3393.18 (1)147.21 + x, −1 + y, z
C9–HB···F5A2.4463.42 (1)173.71 − x, −1/2 + y, − z
C4–HB···F1A2.4213.19 (1)134.61 − x, −1/2 + y, 1 − z
C4–HA···F3A2.4643.10 (1)121.81 − x, 1/2 + y, 1 − z
C10–HC···N52.5843.378 (4)138.1x, −1 + y, z
C5–HB···N72.5553.480 (3)157.51 − x, 1/2 + y, −z
C8–HB···O12.3853.349 (2)168.22 − x, −1/2 + y, 1 − z
Table 3. Selected bond lengths, angles and torsions of the triazene bridge for 13 (Å, °).
Table 3. Selected bond lengths, angles and torsions of the triazene bridge for 13 (Å, °).
Compound123
C1–N51.359 (7)1.353 (3)1.353 (3)
N5–N61.319 (4)1.302 (2)1.304 (3)
N6–N71.296 (5)1.309 (3)1.303 (3)
N7–C61.374 (4)1.351 (2)1.356 (3)
C1–N5–N6110.3 (3)113.3 (2)114.5 (2)
N5–N6–N7111.3 (3)109.8 (2)110.0 (2)
N6–N7–C6110.0 (3)112.9 (2)113.0 (2)
C1–N5–N6–N7173.9 (3)177.7 (2)177.1 (2)
N5–N6–N7–C6175.0 (3)179.9 (2)178.2 (2)
Table 4. Solvent parameters and absorption maxima of the triazenes 13 in polar solvents.
Table 4. Solvent parameters and absorption maxima of the triazenes 13 in polar solvents.
Solventπ*α123
λmax/nmλmax/nmλmax/nm
H2O1.091.17391.0403.0361.5
AcOH0.641.12393.0396.0362.8
MeOH0.600.98395.5394.5364.3
EtOH0.540.86397.5393.7367.5
2-PrOH0.480.76399.0393.7367.0
CH2Cl20.820.13403.5392.0371.7
DMSO1.000405.0
Table 5. Cathodic peak potential (Ep)c and cathodic peak current (Ip)c of the triazenes 13 in CH3CN/0.1 M Bu4PF6 as function of scan rate (potential values vs. ferrocene couple).
Table 5. Cathodic peak potential (Ep)c and cathodic peak current (Ip)c of the triazenes 13 in CH3CN/0.1 M Bu4PF6 as function of scan rate (potential values vs. ferrocene couple).
CompoundScan rate/mV·s−1(Ep)c/mV(Ip)c/µA
120n.d.0.6
50n.d.0.9
100−14702.4
220−14101.8
50−14453.0
100−14753.8
320−13901.2
50−14202.0
100−14302.7

Share and Cite

MDPI and ACS Style

Haslinger, S.; Laus, G.; Kahlenberg, V.; Wurst, K.; Bechtold, T.; Vergeiner, S.; Schottenberger, H. 2-Azidoimidazolium Ions Captured by N-Heterocyclic Carbenes: Azole-Substituted Triazatrimethine Cyanines. Crystals 2016, 6, 40. https://doi.org/10.3390/cryst6040040

AMA Style

Haslinger S, Laus G, Kahlenberg V, Wurst K, Bechtold T, Vergeiner S, Schottenberger H. 2-Azidoimidazolium Ions Captured by N-Heterocyclic Carbenes: Azole-Substituted Triazatrimethine Cyanines. Crystals. 2016; 6(4):40. https://doi.org/10.3390/cryst6040040

Chicago/Turabian Style

Haslinger, Simone, Gerhard Laus, Volker Kahlenberg, Klaus Wurst, Thomas Bechtold, Stefan Vergeiner, and Herwig Schottenberger. 2016. "2-Azidoimidazolium Ions Captured by N-Heterocyclic Carbenes: Azole-Substituted Triazatrimethine Cyanines" Crystals 6, no. 4: 40. https://doi.org/10.3390/cryst6040040

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop